Towards a zero magnetic field environment for ultracold atoms experiments

C. Rogora1    R. Cominotti1    C. Baroni1,2    D. Andreoni1    G. Lamporesi1,3 giacomo.lamporesi@ino.cnr.it    A. Zenesini1,3    G. Ferrari1,3 1Pitaevskii BEC Center, CNR-INO and Dipartimento di Fisica, Università di Trento, 38123 Trento, Italy 2222Institute for Quantum Optics and Quantum Information (IQOQI), Austrian Academy of Sciences, 6020 Innsbruck, Austria 3Trento Institute for Fundamental Physics and Applications, INFN, 38123 Trento, Italy
(July 4, 2024)
Abstract

The minimization of the magnetic field plays a crucial role in ultracold gas research. For instance, the contact interaction dominates all the other energy scales in the zero magnetic field limit, giving rise to novel quantum phases of matter. However, lowering magnetic fields well below the mG level is often challenging in ultracold gas experiments. In this article, we apply Landau-Zener spectroscopy to characterize and reduce the magnetic field on an ultracold gas of sodium atoms to a few tens of μ𝜇\muitalic_μG. The lowest magnetic field achieved here opens the way to observing novel phases of matter with ultracold spinor Bose gases.

I INTRODUCTION

Controlling magnetic fields is critical in many contexts involving fundamental and applied physics experiments and quantum technologies. Often, the performance of a measurement critically depends on the stability against fluctuations of the background magnetic field, as it happens in electron microscopy experiments [1, 2, 3], atom interferometry [4], nuclear magnetic resonance [5], atomic clock experiments [6, 7, 8], and ultracold gases experiments involving coherently-coupled condensate mixtures [9, 10].

In other contexts, such as zero-field nuclear magnetic resonance [11, 12], the constraint relates to the magnitude of the magnetic field, which generally should be minimized and, depending on the type of measurement, needs to be kept below the threshold values. Historically, the measurement of relatively small magnetic fields is achieved using SQUIDs in a cryogenic environment [13], exploiting atomic coherence in room temperature gas cells [14, 15, 16], or atomic spin-alignment [17], with applications in diverse fields, including biomedical imaging [18, 19], metal detection [20, 21], and material characterization [22]. Nowadays, a wide variety of experimental platforms and techniques are available, as described, for instance, in Ref. [23]. In the context of magnetometry, cold gases have been consistently employed as magnetic field sensors [24, 25], in various cases with micrometric scale spatial resolution [26, 27, 28, 29, 30]  exploiting the enhanced sensitivity of spinor condensates to magnetic field inhomogeneities [31].

Finding a way to reduce the magnetic field, control it with high accuracy, and be able to measure such small values would pave the way for the investigation of new physical phenomena, such as the zero-magnetic-field physics of spinor condensates, i.e., condensates with a vector order parameter. Spinor condensates can develop different configurations, and their ground state depends on the strength of the interactions between the internal states and on the strength of the externally applied magnetic field, which removes the degeneracy between different spin states [32]. In the majority of the experiments, the magnetic field is typically large enough that the corresponding magnetic energy splitting between spin states dominates over the spin-dependent interaction energy. In such an experimental configuration, spinorial systems with a nonzero and fixed magnetization can be investigated [33]. The complementary case where the magnetization is set to zero and evolves in the absence of an external magnetic field is also very interesting and yet unexplored, and this work aims at creating the conditions to investigate it experimentally.

A regime where the two energies are comparable has been studied with dipolar chromium gases in the presence of magnetic fields of a few tenths of a mG, thanks to the large magnetic moment of chromium atoms [34].

In ultracold gases of alkali atoms, the contribution of the long-range magnetic dipole interaction is typically small [35], which makes the spin-dependent interaction dominated by the spin-dependent contact interactions. This further reduces the threshold magnetic field below which interesting and unobserved spinor phases are expected with respect to the chromium case [32, 36, 37, 38, 39]. Reducing the amplitude of the magnetic field may allow, for instance, for the observation of fragmentation in sodium condensates without manipulating the energy levels to establish the degeneracy among the Zeeman sublevels [39]. Besides, it is also relevant in the context of dipolar magnetic gases where the magnetic interaction becomes crucial and interesting phenomena can develop such as spontaneous circulation [40].

The strength of the magnetic field below which these novel phases appear is known, in the case of the zero spatial mode approximation [37, 38, 39], to be on the order of a few hundreds of μ𝜇\muitalic_μG. However, this threshold is expected at even lower magnetic fields in the case of spatially extended condensates. Here, the experimental challenge relies on the difficulty of controlling the magnetic field stability at the μ𝜇\muitalic_μG level and reaching and maintaining such small magnetic field values over the whole extension of the atomic sample both during a single experimental sequence and within different runs.

This work presents an experimental technique for minimizing the magnetic field at the 10-μ𝜇\muitalic_μG level. Since it is technically difficult to integrate an external device in the vacuum chamber of ultracold atom experiments to characterize and minimize the field, developing an independent technique for the magnetic field characterization using the trapped atoms as sensors becomes necessary. In particular, we describe an application of Landau-Zener (LZ) spectroscopy [41] over an atomic gas of sodium. Reaching such a low field is possible thanks to the presence of a magnetic shield [42], which demonstrated its efficiency in stabilizing the field in several previous works [43, 44, 45, 46]. The results presented in the following open the way to studying the zero-magnetic-field ground state of an F𝐹Fitalic_F=1 system.

In Sec. II, we describe the experimental platform. Section III contains the theoretical framework and experimental protocols. In Sec. IV we discuss the results, while in Sec. V we report concluding remarks and outlooks.

II Experimental platform

The experimental platform consists of a bosonic gas of 23Na atoms, trapped in the elongated optical trap potential generated by a single far-detuned infrared (1064 nm) laser beam, and therefore equal for all spin components. A thermal sample is loaded in the optical trap and presents a Gaussian spatial density distribution in all three directions. In the optical trap (ωx/2π10similar-tosubscript𝜔x2𝜋10\omega_{\mathrm{x}}/2\pi\sim 10italic_ω start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT / 2 italic_π ∼ 10 Hz, ωy,z/2π1similar-tosubscript𝜔yz2𝜋1\omega_{\mathrm{y,z}}/2\pi\sim 1italic_ω start_POSTSUBSCRIPT roman_y , roman_z end_POSTSUBSCRIPT / 2 italic_π ∼ 1 kHz,), the cloud has an elongated shape with a 1:100 aspect ratio, having the long axis along x𝑥xitalic_x and the short axes along y𝑦yitalic_y and z𝑧zitalic_z.

Refer to caption
Figure 1: Schematics of the core setup. (a) The innermost layer of the magnetic shield (dark gray and transparent) contains a 3D-printed plastic support structure for the coils (red), the coils themselves (blue), and the glass cell (light gray). (b) yz𝑦𝑧yzitalic_y italic_z section of the apparatus.

The sample is prepared inside a magnetic shield, which guarantees the stability of the magnetic field at the  μGtimesabsent𝜇G\text{\,}\mu\mathrm{G}start_ARG end_ARG start_ARG times end_ARG start_ARG italic_μ roman_G end_ARG level [47, 42]. Pairs of magnetic coils inside the innermost layer of the magnetic shield allow for the application of controllable magnetic fields in each of the three spatial directions. Figure 1 schematically represents the apparatus and shows the glass cell, coils, and magnetic shield. The applied magnetic field’s stability is ensured using high-stability current supplies [Stanford research system (SRS) LDC501 for the longitudinal field and Delta Elektronica ES 015-10 with 10:1 current dividers for transverse fields]. Ramping from positive to negative values of the longitudinal field within the same experimental run is achieved using two unipolar current supplies with opposite orientations arranged in parallel. A master SRS is set to the steady drive of 100 mA (similar-to\sim 245 mG), while a second SRS introduces a tunable current in the opposite direction.

III Measurement scheme

Spectroscopic methods are typical solutions to measure a magnetic field. They consist of interrogating an atomic two-level system with constant radiation at different frequencies for a given time and recording the resulting energy spectrum. In such measurements, the Fourier broadening does not represent a limitation at high magnetic fields when the Zeeman splitting (700similar-toabsent700\sim 700∼ 700 kHz/G for atomic 23Na) is much larger than the resonance linewidth. Conversely, when approaching the μ𝜇\muitalic_μG regime, techniques such as radio-frequency spectroscopy are difficult to apply, given the unresolved Zeeman structure, and the difficulty of defining the polarization of the radio-frequency or microwave coupling fields.

An alternative method to characterize the magnetic field around the null value relies on ramps of the magnetic field amplitude and directions, taking advantage of the adiabatic/diabatic dynamics of the atomic spin rotation. At low magnetic field, i.e., when the Larmor frequency is of the order or lower than the velocity of rotation of the magnetic field direction, LZ theory [48, 49, 50] applies, resulting in a powerful tool for the magnetic field characterization.

We developed two protocols based on LZ sweeps on the z𝑧zitalic_z (taken as a quantization axis) component of the field. The first one aims to minimize the magnetic field in the transverse xy𝑥𝑦xyitalic_x italic_y plane and is implemented by ramping (or sweeping) the z𝑧zitalic_z field component from positive to negative finite values. The second protocol involves a ramp with a variable endpoint to find the minimum field along z𝑧zitalic_z.

In the following, we will use the magnetic field nomenclature presented in Tab. 1.

B𝐵Bitalic_B Magnetic field modulus
Bisubscript𝐵𝑖B_{i}italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (i=x,y,z𝑖𝑥𝑦𝑧i=x,y,zitalic_i = italic_x , italic_y , italic_z) Actual magnetic field components
Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT Transverse field
Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT Longitudinal field
Bi,coilssubscript𝐵𝑖coilsB_{i,\mathrm{coils}}italic_B start_POSTSUBSCRIPT italic_i , roman_coils end_POSTSUBSCRIPT Field induced by the i𝑖iitalic_i coil
Bi,rampsubscript𝐵𝑖rampB_{i,\mathrm{ramp}}italic_B start_POSTSUBSCRIPT italic_i , roman_ramp end_POSTSUBSCRIPT Starting Bi,coilssubscript𝐵𝑖coilsB_{i,\mathrm{coils}}italic_B start_POSTSUBSCRIPT italic_i , roman_coils end_POSTSUBSCRIPT of the LZ ramps
Bi,0subscript𝐵𝑖0B_{i,0}italic_B start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT Optimal Bi,coilssubscript𝐵𝑖coilsB_{i,\mathrm{coils}}italic_B start_POSTSUBSCRIPT italic_i , roman_coils end_POSTSUBSCRIPT from fit
Bz,csubscript𝐵𝑧𝑐B_{z,c}italic_B start_POSTSUBSCRIPT italic_z , italic_c end_POSTSUBSCRIPT Field set during condensation
Bz,finsubscript𝐵zfinB_{\mathrm{z,fin}}italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT Final value of longitudinal field protocol
Table 1: Magnetic field nomenclature

III.1 Energy levels and Landau-Zener theory

The hyperfine ground state of sodium has a total angular momentum F=1𝐹1F=1italic_F = 1 with three magnetic sublevels mF=0,±1subscript𝑚𝐹0plus-or-minus1m_{F}=0,\pm 1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 , ± 1 defined as the projection of F𝐹Fitalic_F along the z𝑧zitalic_z-axis. The energy of the three states can be estimated with high accuracy using the Breit-Rabi formula [51] once the magnetic field value is known, and, in the small field limit, a linear dependence on the field (first order Zeeman regime) is expected. Let us consider z𝑧zitalic_z as the quantization axis and the magnetic field with a finite transverse component B=(Bx2+By2)1/2subscript𝐵perpendicular-tosuperscriptsuperscriptsubscript𝐵x2superscriptsubscript𝐵y212B_{\perp}=(B_{\mathrm{x}}^{2}+B_{\mathrm{y}}^{2})^{1/2}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = ( italic_B start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_B start_POSTSUBSCRIPT roman_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. If Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT is linearly ramped with a slope Bz˙˙subscript𝐵z\dot{B_{\mathrm{z}}}over˙ start_ARG italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT end_ARG, from large positive to large negative values passing across zero, the energy of the states as a function of time is the one shown in Fig. 2. At Bz=0subscript𝐵z0B_{\mathrm{z}}=0italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT = 0, the magnetic field is equal to Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT, which introduces an avoided crossing between the states.

Refer to caption
Figure 2: Transverse field and LZ ramps. For each row, the left panels show a schematic view of the transverse magnetic field components (Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT decreases from top to bottom). Central panels show the energy of the three states mF=0,±1subscript𝑚𝐹0plus-or-minus1m_{F}=0,\pm 1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 , ± 1 as a function of time for the corresponding Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. The right panels show the relative populations of the atoms in the three states at the end of the ramp, obtained experimentally and measured through Stern-Gerlach imaging. For the three different values of Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT we obtain full adiabatic transfer (upper panel, large Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT), partial transfer (central panel), and complete diabatic transfer (lower panel, low Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT).

The time-dependent Hamiltonian of the system for the wavefunction {ψ1,ψ0,ψ+1}subscript𝜓1subscript𝜓0subscript𝜓1\{\psi_{-1},\psi_{0},\psi_{+1}\}{ italic_ψ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT } reads, as in [52, 41],

H(t)=gFμB(αtΔ0Δ0Δ0Δαt),𝐻𝑡subscript𝑔𝐹subscript𝜇𝐵matrix𝛼𝑡Δ0Δ0Δ0Δ𝛼𝑡H(t)=g_{F}\mu_{B}\begin{pmatrix}\alpha t&\Delta&0\\ \Delta&0&\Delta\\ 0&\Delta&-\alpha t\\ \end{pmatrix},italic_H ( italic_t ) = italic_g start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_α italic_t end_CELL start_CELL roman_Δ end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL roman_Δ end_CELL start_CELL 0 end_CELL start_CELL roman_Δ end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL roman_Δ end_CELL start_CELL - italic_α italic_t end_CELL end_ROW end_ARG ) , (1)

where gF=1/2subscript𝑔𝐹12g_{F}=-1/2italic_g start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 / 2 is the Landé factor, μBsubscript𝜇𝐵\mu_{B}italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the Bohr magneton, and Δ=B/2Δsubscript𝐵perpendicular-to2\Delta=B_{\perp}/\sqrt{2}roman_Δ = italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / square-root start_ARG 2 end_ARG acts as a coupling between the states and defines the gap in the energy levels shown in \IfBeginWithfig:fig1eq:Eq. (1)\IfBeginWithfig:fig1fig:Fig. 1\IfBeginWithfig:fig1tab:Table 1\IfBeginWithfig:fig1appendix:Appendix 1\IfBeginWithfig:fig1sec:Section 1. Here t=0𝑡0t=0italic_t = 0 is defined as the instant when Bz=0subscript𝐵𝑧0B_{z}=0italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0. The parameter α𝛼\alphaitalic_α originates from the time derivative of the Larmor frequency at the energy levels crossing

Refer to caption
Figure 3: Transverse field minimization protocol. Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT is shown as a function of time. Scheme A (red line) consists in a ramp from Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT to -Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT performed on a thermal cloud, then Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT is brought down to Bz,csubscript𝐵zc-B_{\mathrm{z,c}}- italic_B start_POSTSUBSCRIPT roman_z , roman_c end_POSTSUBSCRIPT in 50 ms before the evaporation in the optical trap. Scheme B (blue line) starts with the condensation at a finite value of the field Bz,csubscript𝐵zcB_{\mathrm{z,c}}italic_B start_POSTSUBSCRIPT roman_z , roman_c end_POSTSUBSCRIPT, then after the evaporation, the field is ramped down to Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT in 50 ms. Here the condensate experiences the field ramp across zero, from Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT to Bz,rampsubscript𝐵zramp-B_{\mathrm{z,ramp}}- italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT. In both cases, the dotted line corresponds to the field crossing the minimum value with a tunable ramp duration. The shaded gray area denotes the time during which the sample is evaporated to obtain a BEC.
α=||B(t)|t(arctanB˙zt|B(t)|)|t=0=|B˙z|,𝛼subscript𝐵𝑡𝑡subscript˙𝐵z𝑡𝐵𝑡𝑡0subscript˙𝐵z\alpha=\left||B(t)|\frac{\partial}{\partial t}\left(\arctan\frac{\dot{B}_{% \mathrm{z}}t}{|B(t)|}\right)\right|_{t=0}=|\dot{B}_{\mathrm{z}}|{\color[rgb]{% 0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke% {0}\pgfsys@color@gray@fill{0},}italic_α = | | italic_B ( italic_t ) | divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ( roman_arctan divide start_ARG over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT italic_t end_ARG start_ARG | italic_B ( italic_t ) | end_ARG ) | start_POSTSUBSCRIPT italic_t = 0 end_POSTSUBSCRIPT = | over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT | , (2)

which relates the Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ramps with the variation of the Larmor precession of the atoms around the applied magnetic field.

Refer to caption
Figure 4: Transverse field determination. (a) Schematic representation of the iterative measurement process; see main text. (b) Relative populations of the three Zeeman levels mF=1,0,+1subscript𝑚𝐹101m_{F}=-1,0,+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 , 0 , + 1 (from top to bottom) for decreasing ramp velocities |B˙z|subscript˙𝐵𝑧{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}|\dot{B}_{z}|}| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | (from left to right). All panels show the experimental data (dots) and the fitted function (solid line) plotted as a function of the transverse field in the x𝑥xitalic_x directions. For each panel, the upper axis shows the field’s value generated by the coils, while the lower axis is shifted according to the fit result. The speed of the ramp is specified in the orange box as well as the value of the compensation field in the y𝑦yitalic_y direction used for each scan. Note the offset of the compensating field in the first column, whose data were acquired with a few weeks delay with respect to the other two data sets.

While the atoms are initially prepared in the mF=1subscript𝑚𝐹1m_{F}=-1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 state at large (as compared to Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT) and positive values of Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT, ramping Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT in time may induce the transfer to different mFsubscript𝑚𝐹m_{F}italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT states according to the analytical model discussed in [41]. In the following, we apply the results of [52] as we measure the mFsubscript𝑚𝐹m_{F}italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT population distribution at long times after the inversion of Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT:

P1subscript𝑃1\displaystyle P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =p2,absentsuperscript𝑝2\displaystyle=p^{2},= italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (3)
P0subscript𝑃0\displaystyle P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =2p(1p),absent2𝑝1𝑝\displaystyle=2p(1-p),= 2 italic_p ( 1 - italic_p ) , (4)
P1subscript𝑃1\displaystyle P_{-1}italic_P start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT =(1p)2,absentsuperscript1𝑝2\displaystyle=(1-p)^{2},= ( 1 - italic_p ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (5)

with

p=e2π|gF|μBΔ22|Bz˙|𝑝superscript𝑒2𝜋subscript𝑔𝐹subscript𝜇𝐵superscriptΔ22Planck-constant-over-2-pi˙subscript𝐵𝑧p=e^{-2\pi|g_{F}|\mu_{B}\frac{\Delta^{2}}{{\color[rgb]{0,0,0}\definecolor[% named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}2}\hbar{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}|\dot{B_{z}}|}}}italic_p = italic_e start_POSTSUPERSCRIPT - 2 italic_π | italic_g start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT | italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT divide start_ARG roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 roman_ℏ | over˙ start_ARG italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG | end_ARG end_POSTSUPERSCRIPT (6)

being the LZ transfer probability. From Eq. (3) and Eq. (6), one finds that a complete diabatic transfer of the population from the mF=1subscript𝑚𝐹1m_{F}=-1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1 to the mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1 state takes place when p1𝑝1p\approx 1italic_p ≈ 1 (large ramp speed |B˙z|subscript˙𝐵𝑧{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}|\dot{B}_{z}|}| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | or small gap ΔΔ\Deltaroman_Δ), while adiabaticity is preserved for small |B˙z|subscript˙𝐵𝑧{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}|\dot{B}_{z}|}| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | and large ΔΔ\Deltaroman_Δ. This results from implementing the adiabatic condition to the spin dynamics in a rotating magnetic field. In other words, adiabaticity is fulfilled when the variation of the magnetic field direction is smaller than the Larmor frequency, i.e., |Bz˙/B||gFμBB/|much-less-than˙subscript𝐵z𝐵subscript𝑔𝐹subscript𝜇𝐵𝐵Planck-constant-over-2-pi{|\dot{B_{\mathrm{z}}}}/{B}|\ll|{g_{F}\mu_{B}B}/{\hbar}|| over˙ start_ARG italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT end_ARG / italic_B | ≪ | italic_g start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_B / roman_ℏ |. For instance, the probability of the transition to mF=1subscript𝑚𝐹1m_{F}=1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 1 has a Gaussian distribution with RMS width

σB=B˙z2π|gF|μB,subscript𝜎𝐵Planck-constant-over-2-pisubscript˙𝐵z2𝜋subscript𝑔𝐹subscript𝜇𝐵\sigma_{B}=\sqrt{\frac{\hbar\dot{B}_{\mathrm{z}}}{2\pi|g_{F}|\mu_{B}}},italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG roman_ℏ over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π | italic_g start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT | italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG end_ARG , (7)

which is the width used in the following.

III.2 Experimental protocols

III.2.1 Transverse field minimization

The goal is to find the optimal current values for each coil to compensate for the residual transverse magnetic field (which is not screened by the magnetic shield or due to the shield’s permanent magnetization).

The protocol to minimize the transverse field amplitude Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT starts with a thermal atomic sample by setting the values of the currents in the coils for the transverse directions, generating the fields Bx,coilssubscript𝐵xcoilsB_{\mathrm{x,coils}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT and By,coilssubscript𝐵ycoilsB_{\mathrm{y,coils}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT. In the following, we discuss the two experimental schemes depicted in \IfBeginWithfig:fig3eq:Eq. (3)\IfBeginWithfig:fig3fig:Fig. 3\IfBeginWithfig:fig3tab:Table 3\IfBeginWithfig:fig3appendix:Appendix 3\IfBeginWithfig:fig3sec:Section 3.

Scheme A (red line) consists in changing Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT from Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT to Bz,rampsubscript𝐵zramp-B_{\mathrm{z,ramp}}- italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT with a linear ramp of variable duration ΔtΔsuperscript𝑡\Delta t^{*}roman_Δ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. After this first ramp, Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT is reduced to Bz,c130subscript𝐵zc130-B_{\mathrm{z,c}}\approx-130- italic_B start_POSTSUBSCRIPT roman_z , roman_c end_POSTSUBSCRIPT ≈ - 130 mG Bz,rampmuch-less-thanabsentsubscript𝐵zramp\ll-B_{\mathrm{z,ramp}}≪ - italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT in 50 ms. Then, the sample is evaporatively cooled to Bose condensation by reducing the dipole trapping beam intensity. To image the spin state, after switching off the trapping potential, a vertical magnetic field gradient of 8 G/cm is applied to separate the three states in a Stern-Gerlach-like scheme. A simultaneous absorption image of the three states is made after a time-of-flight of 18 ms. Bose condensing the sample before imaging favors the spatial spin resolution of the Stern-Gerlach imaging, given the relatively small amount of the applied magnetic field gradient and ballistic expansion time.

To reduce the sensitivity to magnetic field inhomogeneities using an atomic sample with reduced spatial extension, in Scheme B (blue line) the order between LZ ramp and evaporation is reversed. The atomic sample is first cooled below the critical temperature for condensation at Bz,csubscript𝐵zcabsentB_{\mathrm{z,c}}\approxitalic_B start_POSTSUBSCRIPT roman_z , roman_c end_POSTSUBSCRIPT ≈ 130 mG, and then the field is decreased to Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT to apply the LZ ramp from Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT to Bz,rampsubscript𝐵zramp-B_{\mathrm{z,ramp}}- italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT. By doing so, the spatial extent of the condensed atomic sample is considerably smaller during the LZ ramp, suppressing the contributions from the magnetic field’s spatial inhomogeneities.

The ramp duration ΔtΔsuperscript𝑡\Delta t^{*}roman_Δ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT should be much longer than the coils time constant τ=L/R𝜏𝐿𝑅similar-to-or-equalsabsent\tau=L/R\simeqitalic_τ = italic_L / italic_R ≃ 0.5 ms, where L𝐿Litalic_L and R𝑅Ritalic_R are the inductance and the resistance of the coils, respectively, and much shorter than the sample lifetime in the trap (τBECsimilar-to-or-equalssubscript𝜏BECabsent\tau_{\mathrm{BEC}}\simeqitalic_τ start_POSTSUBSCRIPT roman_BEC end_POSTSUBSCRIPT ≃ 1 s, τthmuch-greater-thansubscript𝜏thabsent\tau_{\mathrm{th}}\ggitalic_τ start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT ≫ 1 s). The value Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT was chosen depending on the ramp duration, from a maximum of 1.5 mG to a minimum of 750 μ𝜇\muitalic_μG, with |Bz˙|˙subscript𝐵z|\dot{B_{\mathrm{z}}}|| over˙ start_ARG italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT end_ARG | ranging from a maximum of 3 G/s to a minimum of 3.75 mG/s.

The values Bx,0subscript𝐵x0B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT and By,0subscript𝐵y0B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT that best compensate for the residual transverse field are the ones giving the maximal transfer to the mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1 state. They are found through an iterative procedure, as sketched in \IfBeginWithfig:fig4eq:Eq. (4)\IfBeginWithfig:fig4fig:Fig. 4\IfBeginWithfig:fig4tab:Table 4\IfBeginWithfig:fig4appendix:Appendix 4\IfBeginWithfig:fig4sec:Section 4a. The first iteration is a scan in Bx,coilssubscript𝐵xcoilsB_{\mathrm{x,coils}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT setting |B˙z|=300subscript˙𝐵z300|\dot{B}_{\mathrm{z}}|=300| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT | = 300 mG/s and By,coils=0subscript𝐵ycoils0B_{\mathrm{y,coils}}=0italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT = 0. We obtain the value Bx,0subscript𝐵x0B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT for which we observe maximum transfer in mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1. Then, we set Bx,coils=Bx,0subscript𝐵xcoilssubscript𝐵x0B_{\mathrm{x,coils}}=B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT and we perform a scan in By,coilssubscript𝐵ycoilsB_{\mathrm{y,coils}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT obtaining By,0subscript𝐵y0B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT for which we have the maximal transfer of the population in the state mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1. We repeat this procedure by slowing down the ramp in Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT at each iteration. In this way, the P+1subscript𝑃1P_{+1}italic_P start_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT distribution shrinks, increasing our sensitivity to determine the field value that best compensates for the residual one. The first iteration starts finding the value Bx,0subscript𝐵x0B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT with a scan in Bx,coilssubscript𝐵xcoilsB_{\mathrm{x,coils}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT at fixed By,coils=0subscript𝐵ycoils0B_{\mathrm{y,coils}}=0italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT = 0, with a given ramp speed |B˙z|=300subscript˙𝐵z300|\dot{B}_{\mathrm{z}}|=300| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT | = 300 mG/s. Then, at fixed Bx,coils=Bx,0subscript𝐵xcoilssubscript𝐵x0B_{\mathrm{x,coils}}=B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT we perform a scan in By,coilssubscript𝐵ycoilsB_{\mathrm{y,coils}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT obtaining By,0subscript𝐵y0B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT.

Refer to caption
Figure 5: Longitudinal field protocol. Representation of the field in z𝑧zitalic_z and y𝑦yitalic_y directions. For the field in the longitudinal direction, the protocol is analogous to Scheme B with the only difference that the speed of the ramp is fixed to be B˙z=15subscript˙𝐵z15\dot{B}_{\mathrm{z}}=15over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT = 15mG/s while the final value of the ramp Bz,finsubscript𝐵zfinB_{\mathrm{z,fin}}italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT is varied. The field in the y𝑦yitalic_y-direction is fixed to the optimal value By,0subscript𝐵y0B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT previously determined, it is adiabatically ramped to a finite value and then back o By,0subscript𝐵y0B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT for the imaging procedure. The field along x𝑥xitalic_x is not shown in the figure, as it is kept at a constant value Bx,0subscript𝐵x0B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT for the whole duration of the experiment.

Figure 4b presents the three hyperfine relative populations, n1subscript𝑛1n_{-1}italic_n start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, n0subscript𝑛0n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, n+1subscript𝑛1n_{+1}italic_n start_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT, for three subsequent scans of the magnetic field Bx,coilssubscript𝐵xcoilsB_{\mathrm{x,coils}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT, with decreasing value of |B˙z|subscript˙𝐵z|\dot{B}_{\mathrm{z}}|| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT | and By,coilssubscript𝐵ycoilsB_{\mathrm{y,coils}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT set to the value determined in the previous iteration step. The experimental data were fitted to Eq. (3-5) from which we extract the center Bx,0subscript𝐵x0B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT, the maximum transfer P1,maxsubscript𝑃1maxP_{1,\mathrm{max}}italic_P start_POSTSUBSCRIPT 1 , roman_max end_POSTSUBSCRIPT, and the width σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT of the transfer peak. At first, the center of the observed structures corresponds to Bx=0subscript𝐵x0B_{\mathrm{x}}=0italic_B start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT = 0 and this allows us to determine Bx,coilssubscript𝐵xcoilsB_{\mathrm{x,coils}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT to compensate any residual field along x𝑥xitalic_x. The residual field in y𝑦yitalic_y depends on the maximum in the transferred population P1,maxsubscript𝑃1maxP_{1,\mathrm{max}}italic_P start_POSTSUBSCRIPT 1 , roman_max end_POSTSUBSCRIPT as

By=σB2ln(P1,max),subscript𝐵𝑦subscript𝜎𝐵2subscript𝑃1maxB_{y}=\sigma_{B}\sqrt{-2\ln(P_{1,\mathrm{max}})}\,,italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT square-root start_ARG - 2 roman_ln ( italic_P start_POSTSUBSCRIPT 1 , roman_max end_POSTSUBSCRIPT ) end_ARG , (8)

which is minimized after each Bxsubscript𝐵xB_{\mathrm{x}}italic_B start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT scan (scan in By,coilssubscript𝐵ycoilsB_{\mathrm{y,coils}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT are not shown in figure). As expected, slower ramps lead to smaller σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT (as highlighted by the white windows, which always mark a region of 200 μ𝜇\muitalic_μG), and consequently to increase the precision at which Bx=0subscript𝐵x0B_{\mathrm{x}}=0italic_B start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT = 0 (and Bysubscript𝐵yB_{\mathrm{y}}italic_B start_POSTSUBSCRIPT roman_y end_POSTSUBSCRIPT) is determined.

III.2.2 Longitudinal field

Refer to caption
Figure 6: Example of the longitudinal field characterization. Each panel shows the relative population for the three states as a function of Bz,finsubscript𝐵zfinB_{\mathrm{z,fin}}italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT. The dots are experimental data obtained by averaging different experimental realizations, with standard deviation smaller than the marker size, while dotted line is the time dependent model. The value Bz,0subscript𝐵z0B_{\mathrm{z,0}}italic_B start_POSTSUBSCRIPT roman_z , 0 end_POSTSUBSCRIPT compensating for the residual Bzsubscript𝐵𝑧B_{z}italic_B start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT field is indicated by the dotted grey line.

The protocol presented in the previous section allows for minimizing the transverse magnetic field. The best-obtained values are then used as transverse fields setting Bx,coils=Bx,0subscript𝐵xcoilssubscript𝐵x0B_{\mathrm{x,coils}}=B_{\mathrm{x,0}}italic_B start_POSTSUBSCRIPT roman_x , roman_coils end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT roman_x , 0 end_POSTSUBSCRIPT and By,coils=By,0subscript𝐵ycoilssubscript𝐵y0B_{\mathrm{y,coils}}=B_{\mathrm{y,0}}italic_B start_POSTSUBSCRIPT roman_y , roman_coils end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT roman_y , 0 end_POSTSUBSCRIPT for the characterization of the residual longitudinal field Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT.

Here, we introduce the procedure to characterize the residual longitudinal field Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT. The gas is first evaporated to obtain a BEC at Bz,c=130subscript𝐵zc130B_{\mathrm{z,c}}=130italic_B start_POSTSUBSCRIPT roman_z , roman_c end_POSTSUBSCRIPT = 130 mG, then Bz,coilssubscript𝐵zcoilsB_{\mathrm{z,coils}}italic_B start_POSTSUBSCRIPT roman_z , roman_coils end_POSTSUBSCRIPT is ramped from a positive value Bz,rampsubscript𝐵zrampB_{\mathrm{z,ramp}}italic_B start_POSTSUBSCRIPT roman_z , roman_ramp end_POSTSUBSCRIPT down to a variable one Bz,finsubscript𝐵zfinB_{\mathrm{z,fin}}italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT with constant ramp B˙z=0.39subscript˙𝐵z0.39\dot{B}_{\mathrm{z}}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{% rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}-0.39}over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT = - 0.39 G/s. As the Stern-Gerlach imaging is implemented at positive Bzsubscript𝐵zB_{\mathrm{z}}italic_B start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT, we constrain the diabatic spin dynamics to the decreasing ramp on Bz,coilssubscript𝐵zcoilsB_{\mathrm{z,coils}}italic_B start_POSTSUBSCRIPT roman_z , roman_coils end_POSTSUBSCRIPT by raising the transverse field to a finite value after the end of the Bz,coilssubscript𝐵zcoilsB_{\mathrm{z,coils}}italic_B start_POSTSUBSCRIPT roman_z , roman_coils end_POSTSUBSCRIPT ramp, see Fig. 5. If Bz,fin>0subscript𝐵zfin0B_{\mathrm{z,fin}}>0italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT > 0, the LZ transfer does not take place since the ramp is interrupted before the zero crossing on Bz,coilssubscript𝐵zcoilsB_{\mathrm{z,coils}}italic_B start_POSTSUBSCRIPT roman_z , roman_coils end_POSTSUBSCRIPT hence leaving the atoms in mF=1subscript𝑚𝐹1m_{F}=-1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 1. If Bz,coils<0subscript𝐵zcoils0B_{\mathrm{z,coils}}<0italic_B start_POSTSUBSCRIPT roman_z , roman_coils end_POSTSUBSCRIPT < 0, on the other hand, the transfer to mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1 takes place. An example of such a scan is given in Fig. 6 where the relative populations in the three states are represented as a function of the value Bz,finsubscript𝐵zfinB_{\mathrm{z,fin}}italic_B start_POSTSUBSCRIPT roman_z , roman_fin end_POSTSUBSCRIPT. We verified that using a slower ramp does not affect the experimental result. The line in the plot is a fit based on the time dependent model reported in Ref.[41]. The extracted central position allows to determine Bz,0subscript𝐵𝑧0B_{z,0}italic_B start_POSTSUBSCRIPT italic_z , 0 end_POSTSUBSCRIPT. Note that the typical oscillations of the LZ process are not visible here due to the large population transfer, which suppresses interference process between different states.

IV RESULTS

Refer to caption
Figure 7: a) Width σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT of LZ transfer probability as a function of the ramp speed. The dashed black line is the theoretical prediction from LZ theory; the symbols are the experimental results using scheme A (red symbols) or scheme B (blue symbols). Empty (filled) symbols are obtained without (with) compensation for the residual gradient, respectively. The error bars have been estimated considering the fluctuations of different datasets. The two insets are examples of the population in the state mF=+1subscript𝑚𝐹1m_{F}=+1italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = + 1 obtained at a ramp velocity of 3.75 mG/s with (left inset) and without (right inset) the compensation of the gradient. b) Transverse residual magnetic field Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT extracted from the population fits as a function of the inverse of the ramp velocities. Note that the error bars for the data obtained compensating the magnetic field inhomogeneities are too small to be seen in the plot.

From the fits of the experimental data, as the examples shown in Fig. 4, it is possible to extract both the width of the transfer peak σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT and, from the contrast of the LZ population transfer, the residual transverse magnetic field Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. Figure 7a shows the measured width σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT of the transfer peak as a function of |B˙z|subscript˙𝐵z|\dot{B}_{\mathrm{z}}|| over˙ start_ARG italic_B end_ARG start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT | for both protocols (red symbols for A and blue for B). Each point was obtained by averaging the widths from different experimental runs with the same ramp speed. The experimental values of the width are consistent with σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT predicted from the LZ theory (dashed line in the figure).

For instance, the flattening behavior observed in Fig.7 around σB=20μsubscript𝜎𝐵20𝜇\sigma_{B}=20\,\muitalic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 20 italic_μG, and B=25μsubscript𝐵perpendicular-to25𝜇B_{\perp}=25\,\muitalic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 25 italic_μG, both for thermal and condensed samples, suggests the possible role of magnetic field inhomogeneities, in particular along the long axis of the condensate. We add a magnetic field gradient along the axial direction of the sample to further reduce σBsubscript𝜎𝐵\sigma_{B}italic_σ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT and Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT, as shown by the filled dots in \IfBeginWithfig:fig7eq:Eq. (7)\IfBeginWithfig:fig7fig:Fig. 7\IfBeginWithfig:fig7tab:Table 7\IfBeginWithfig:fig7appendix:Appendix 7\IfBeginWithfig:fig7sec:Section 7a and \IfBeginWithfig:fig7eq:Eq. (7)\IfBeginWithfig:fig7fig:Fig. 7\IfBeginWithfig:fig7tab:Table 7\IfBeginWithfig:fig7appendix:Appendix 7\IfBeginWithfig:fig7sec:Section 7b.

From the maximum transmission probability P1,maxsubscript𝑃1maxP_{\mathrm{1,max}}italic_P start_POSTSUBSCRIPT 1 , roman_max end_POSTSUBSCRIPT obtained from the different scans, we can calculate Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT by using Eq. (8) both for measurements without (empty symbols) and with (filled triangles) the gradient compensation, as shown in \IfBeginWithfig:fig7eq:Eq. (7)\IfBeginWithfig:fig7fig:Fig. 7\IfBeginWithfig:fig7tab:Table 7\IfBeginWithfig:fig7appendix:Appendix 7\IfBeginWithfig:fig7sec:Section 7b. The smallest width observed allows us to determine the minimal residual transverse field at (14±2)μplus-or-minus142𝜇(14\pm 2)\,\mu( 14 ± 2 ) italic_μG, a value that could be limited by the noise of the current supplies driving the compensation coils. Also, it is worth mentioning that the condition for the residual field compensation has generally been stable for several weeks. Still, we did observe jumps in the compensation field at the level of hundreds of μ𝜇\muitalic_μG (a few events over six months of measurements), which we could not clearly attribute to technical circumstances.

The longitudinal field was minimized by following the protocol explained in Sec. III.2.2. Bz,0subscript𝐵z0B_{\mathrm{z,0}}italic_B start_POSTSUBSCRIPT roman_z , 0 end_POSTSUBSCRIPT is chosen as the center of the time-dependent model fitted to the data shown in Fig. 6, Bz,0=(323±10)μsubscript𝐵z0plus-or-minus32310𝜇B_{\mathrm{z,0}}=(-323\pm 10)\,\muitalic_B start_POSTSUBSCRIPT roman_z , 0 end_POSTSUBSCRIPT = ( - 323 ± 10 ) italic_μG.

By combining the minimal transverse and longitudinal field results, we estimate we can reach a minimal field modulus of (18±5)μplus-or-minus185𝜇(18\pm 5)\,\mu( 18 ± 5 ) italic_μG, which complies with the conditions for observing a nematic phase in an elongated 23Na condensate.

V CONCLUSIONS AND OUTLOOK

In this paper, we present a technique to characterize and compensate magnetic fields at the level of 10 μ𝜇\muitalic_μG in experiments with ultracold atomic gases. The method is based on monitoring the Zeeman populations in diabatic atomic spin rotation. The simulations based on the LZ dynamics for a three-level system reproduced the experimental data well. These results pave the way to studying the unexplored scenario of condensation in zero magnetic fields in extended spinor gases, when spin interactions may become dominant over all other contributions to the Hamiltonian. For instance, the ground state of an F=1 spinor condensate, characterized by antiferromagnetic interactions, develops an order parameter with a nematic character, as observed in the single spatial mode approximation [37, 38], but its superfluid properties were not explored so far.

Acknowledgements

We acknowledge funding from Provincia Autonoma di Trento, from the European Union’s Horizon 2020 research and innovation Programme through the STAQS project of QuantERA II (Grant Agreements No. 101017733 and No. 101017733), from the project DYNAMITE QUANTERA2-00056 funded by the Ministry of University and Research through the ERANET COFUND QuantERA II – 2021 call and co-funded by the European Union, and from the European Union - Next Generation EU through PNRR MUR project PE0000023-NQSTI. This work was supported by Q@TN, the joint lab between the University of Trento, FBK - Fondazione Bruno Kessler, INFN - National Institute for Nuclear Physics and CNR - National Research Council.

References

  • Mulvey [1962] T. Mulvey, Origins and historical development of the electron microscope, British Journal of Applied Physics 13, 197 (1962).
  • Ruska [1987] E. Ruska, The development of the electron microscope and of electron microscopy, Rev. Mod. Phys. 59, 627 (1987).
  • Krivanek et al. [2008] O. Krivanek, G. Corbin, N. Dellby, B. Elston, R. Keyse, M. F. Murfitt, C. Own, Z. Szilagyi, and J. Woodruff, An electron microscope for the aberration-corrected era, Ultramicroscopy 108, 179 (2008).
  • [4] E. Rasel, W. Schleich, and S. Wölk, eds., Atom interferometry and its applications (IOS, Amsterdam; SIF, Bologna 2019) pp. 345–392, Proceedings of the International School of Physics “Enrico Fermi”, Course CXCVII, Varenna, 8-13 July 2016.
  • Mansfield and Chapman [1987] P. Mansfield and B. Chapman, Multishield active magnetic screening of coil structures in NMR, Journal of Magnetic Resonance 72, 211 (1987).
  • Clairon et al. [1991] A. Clairon, C. Salomon, S. Guellati, and W. D. Phillips, Ramsey resonance in a zacharias fountain, Europhysics Letters 16, 165 (1991).
  • Gibble and Chu [1993] K. Gibble and S. Chu, Laser-cooled cs frequency standard and a measurement of the frequency shift due to ultracold collisions, Phys. Rev. Lett. 70, 1771 (1993).
  • Ludlow et al. [2015] A. D. Ludlow, M. M. Boyd, J. Ye, E. Peik, and P. O. Schmidt, Optical atomic clocks, Rev. Mod. Phys. 87, 637 (2015).
  • Nicklas et al. [2015] E. Nicklas, M. Karl, M. Höfer, A. Johnson, W. Muessel, H. Strobel, J. Tomkovič, T. Gasenzer, and M. K. Oberthaler, Observation of scaling in the dynamics of a strongly quenched quantum gas, Phys. Rev. Lett. 115, 245301 (2015).
  • Cominotti et al. [2024] R. Cominotti, C. Rogora, A. Zenesini, G. Lamporesi, and G. Ferrari, Ultracold atomic spin mixtures in ultrastable magnetic field environments, Europhysics Letters 146, 45001 (2024).
  • Ledbetter et al. [2008] M. P. Ledbetter, I. M. Savukov, D. Budker, V. Shah, S. Knappe, J. Kitching, D. J. Michalak, S. Xu, and A. Pines, Zero-field remote detection of nmr with a microfabricated atomic magnetometer, Proceedings of the National Academy of Sciences 105, 2286 (2008).
  • Tayler et al. [2017] M. C. D. Tayler, T. Theis, T. F. Sjolander, J. W. Blanchard, A. Kentner, S. Pustelny, A. Pines, and D. Budker, Invited Review Article: Instrumentation for nuclear magnetic resonance in zero and ultralow magnetic field, Review of Scientific Instruments 88, 091101 (2017).
  • Romani et al. [1982] G. L. Romani, S. J. Williamson, and L. Kaufman, Biomagnetic instrumentation, Review of Scientific Instruments 53, 1815 (1982).
  • Dupont-Roc et al. [1969] J. Dupont-Roc, S. Haroche, and C. Cohen-Tannoudji, Detection of very weak magnetic fields (109Gsuperscript109G10^{-9}\,\mathrm{G}10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT roman_G) by Rb87superscriptRb87{}^{87}\mathrm{Rb}start_FLOATSUPERSCRIPT 87 end_FLOATSUPERSCRIPT roman_Rb zero-field level crossing resonances, Physics Letters A 28, 638 (1969).
  • Budker et al. [2002] D. Budker, W. Gawlik, D. F. Kimball, S. M. Rochester, V. V. Yashchuk, and A. Weis, Resonant nonlinear magneto-optical effects in atoms, Rev. Mod. Phys. 74, 1153 (2002).
  • Budker and Romalis [2007] D. Budker and M. Romalis, Optical magnetometry, Nature Physics 3, 227 (2007).
  • Meraki et al. [2023] A. Meraki, L. Elson, N. Ho, A. Akbar, M. Koźbiał, J. Kołodyński, and K. Jensen, Zero-field optical magnetometer based on spin alignment, Phys. Rev. A 108, 062610 (2023).
  • Brookes et al. [2022] M. Brookes, J. Leggett, M. Rea, R. Hill, N. Holmes, E. Boto, and R. Bowtell, Magnetoencephalography with optically pumped magnetometers (OPM-MEG): the next generation of functional neuroimaging, Trends Neuroscience 8, 621 (2022).
  • Sander et al. [2012] T. Sander, J. Preusser, R. Mhaskar, J. Kitching, L. Trahms, and S. Knappe, Magnetoencephalography with a chip-scale atomic magnetometer, Biomed Opt Express 3, 981 (2012).
  • Bevington et al. [2019] P. Bevington, R. Gartman, and W. Chalupczak, Enhanced material defect imaging with a radio-frequency atomic magnetometer, Journal of Applied Physics 125, 094503 (2019).
  • Rushton et al. [2022] L. Rushton, T. Pyragius, A. Meraki, L. Elson, and K. Jensen, Unshielded portable optically pumped magnetometer for the remote detection of conductive objects using eddy current measurements, Review of Scientific Instruments 93, 125103 (2022).
  • Romalis and Dang [2011] M. V. Romalis and H. B. Dang, Atomic magnetometers for materials characterization, Materials Today 14, 258 (2011).
  • Mitchell and Palacios Alvarez [2020] M. W. Mitchell and S. Palacios Alvarez, Colloquium: Quantum limits to the energy resolution of magnetic field sensors, Rev. Mod. Phys. 92, 021001 (2020).
  • Isayama et al. [1999] T. Isayama, Y. Takahashi, N. Tanaka, K. Toyoda, K. Ishikawa, and T. Yabuzaki, Observation of larmor spin precession of laser-cooled rb atoms via paramagnetic faraday rotation, Phys. Rev. A 59, 4836 (1999).
  • Cohen et al. [2019] Y. Cohen, K. Jadeja, S. Sula, M. Venturelli, C. Deans, L. Marmugi, and F. Renzoni, A cold atom radio-frequency magnetometer, Applied Physics Letters 114, 073505 (2019).
  • Higbie et al. [2005] J. M. Higbie, L. E. Sadler, S. Inouye, A. P. Chikkatur, S. R. Leslie, K. L. Moore, V. Savalli, and D. M. Stamper-Kurn, Direct nondestructive imaging of magnetization in a spin-1 bose-einstein gas, Phys. Rev. Lett. 95, 050401 (2005).
  • Wildermuth et al. [2006] S. Wildermuth, S. Hofferberth, I. Lesanovsky, S. Groth, P. Krüger, J. Schmiedmayer, and I. Bar-Joseph, Sensing electric and magnetic fields with Bose-Einstein condensates, Applied Physics Letters 88, 264103 (2006).
  • Vengalattore et al. [2007] M. Vengalattore, J. M. Higbie, S. R. Leslie, J. Guzman, L. E. Sadler, and D. M. Stamper-Kurn, High-resolution magnetometry with a spinor bose-einstein condensate, Phys. Rev. Lett. 98, 200801 (2007).
  • Elíasson et al. [2019] O. Elíasson, R. Heck, J. S. Laustsen, M. Napolitano, R. Müller, M. G. Bason, J. J. Arlt, and J. F. Sherson, Spatially-selective in situ magnetometry of ultracold atomic clouds, Journal of Physics B: Atomic, Molecular and Optical Physics 52, 075003 (2019).
  • Yang et al. [2020] F. Yang, S. F. Taylor, S. D. Edkins, J. C. Palmstrom, I. R. Fisher, and B. L. Lev, Nematic transitions in iron pnictide superconductors imaged with a quantum gas, Nature Physics 16, 514 (2020).
  • Witkowski et al. [2013] M. Witkowski, R. Gartman, B. Nagórny, M. Piotrowski, M. Płodzień, K. Sacha, J. Szczepkowski, J. Zachorowski, M. Zawada, and W. Gawlik, Matter-wave interference versus spontaneous pattern formation in spinor bose-einstein condensates, Phys. Rev. A 88, 025602 (2013).
  • Stamper-Kurn and Ueda [2013] D. M. Stamper-Kurn and M. Ueda, Spinor Bose gases: Symmetries, magnetism, and quantum dynamics, Rev. Mod. Phys. 85, 1191 (2013).
  • Chomaz et al. [2022] L. Chomaz, I. Ferrier-Barbut, F. Ferlaino, B. Laburthe-Tolra, B. L. Lev, and T. Pfau, Dipolar physics: a review of experiments with magnetic quantum gases, Reports on Progress in Physics 86, 026401 (2022).
  • Pasquiou et al. [2011] B. Pasquiou, E. Maréchal, G. Bismut, P. Pedri, L. Vernac, O. Gorceix, and B. Laburthe-Tolra, Spontaneous Demagnetization of a Dipolar Spinor Bose Gas in an Ultralow Magnetic Field, Phys. Rev. Lett. 106, 255303 (2011).
  • Fattori et al. [2008] M. Fattori, G. Roati, B. Deissler, C. D’Errico, M. Zaccanti, M. Jona-Lasinio, L. Santos, M. Inguscio, and G. Modugno, Magnetic dipolar interaction in a bose-einstein condensate atomic interferometer, Phys. Rev. Lett. 101, 190405 (2008).
  • Jiménez-García et al. [2019] K. Jiménez-García, A. Invernizzi, B. Evrard, C. Frapolli, J. Dalibard, and F. Gerbier, Spontaneous formation and relaxation of spin domains in antiferromagnetic spin-1 condensates, Nature Communication 10, 1422 (2019).
  • Hamley et al. [2012] C. D. Hamley, C. S. Gerving, T. M. Hoang, E. M. Bookjans, and M. S. Chapman, Spin-nematic squeezed vacuum in a quantum gas, Nature Physics 8, 305 (2012).
  • Zibold et al. [2016] T. Zibold, V. Corre, C. Frapolli, A. Invernizzi, J. Dalibard, and F. Gerbier, Spin-nematic order in antiferromagnetic spinor condensates, Phys. Rev. A 93, 023614 (2016).
  • Evrard et al. [2021] B. Evrard, A. Qu, J. Dalibard, and F. Gerbier, Observation of fragmentation of a spinor bose-einstein condensate, Science 373, 1340 (2021).
  • Kawaguchi et al. [2006] Y. Kawaguchi, H. Saito, and M. Ueda, Spontaneous circulation in ground-state spinor dipolar bose-einstein condensates, Phys. Rev. Lett. 97, 130404 (2006).
  • Band and Avishai [2019] Y. B. Band and Y. Avishai, Three-level Landau-Zener dynamics, Phys. Rev. A 99, 032112 (2019).
  • Farolfi et al. [2019] A. Farolfi, D. Trypogeorgos, G. Colzi, E. Fava, G. Lamporesi, and G. Ferrari, Design and characterization of a compact magnetic shield for ultracold atomic gas experiments, Review of Scientific Instruments 90, 115114 (2019).
  • Farolfi et al. [2021a] A. Farolfi, A. Zenesini, D. Trypogeorgos, C. Mordini, A. Gallemí, A. Roy, A. Recati, G. Lamporesi, and G. Ferrari, Quantum-torque-induced breaking of magnetic interfaces in ultracold gases, Nat. Phys. 17, 1359 (2021a).
  • Farolfi et al. [2021b] A. Farolfi, A. Zenesini, R. Cominotti, D. Trypogeorgos, A. Recati, G. Lamporesi, and G. Ferrari, Manipulation of an elongated internal Josephson junction of bosonic atoms, Phys. Rev. A 104, 023326 (2021b).
  • Cominotti et al. [2023] R. Cominotti, A. Berti, C. Dulin, C. Rogora, G. Lamporesi, I. Carusotto, A. Recati, A. Zenesini, and G. Ferrari, Ferromagnetism in an Extended Coherently Coupled Atomic Superfluid, Phys. Rev. X 13, 021037 (2023).
  • Zenesini et al. [2024] A. Zenesini, A. Berti, R. Cominotti, C. Rogora, I. G. Moss, T. P. Billam, I. Carusotto, G. Lamporesi, A. Recati, and G. Ferrari, False vacuum decay via bubble formation in ferromagnetic superfluids, Nature Physics 20, 558 (2024).
  • Colzi et al. [2018] G. Colzi, E. Fava, M. Barbiero, C. Mordini, G. Lamporesi, and G. Ferrari, Production of large Bose-Einstein condensates in a magnetic-shield-compatible hybrid trap, Phys. Rev. A 97, 053625 (2018).
  • Landau [1932] L. Landau, Zur theorie der energieubertragung ii, Zeitschrift fur Physik (Sowjetunion) 2, 46 (1932).
  • Zener [1932] C. Zener, Non-adiabatic crossing of energy levels, Proc. R. Soc. Lond. A137, 696–702 (1932).
  • Kofman et al. [2023] P. O. Kofman, O. V. Ivakhnenko, S. N. Shevchenko, and F. Nori, Majorana’s approach to nonadiabatic transitions validates the adiabatic-impulse approximation, Scientific Reports 13, 5053 (2023).
  • Breit and Rabi [1931] G. Breit and I. I. Rabi, Measurement of nuclear spin, Phys. Rev. 38, 2082 (1931).
  • Carroll and Hioe [1986] C. E. Carroll and F. T. Hioe, Transition probabilities for the three-level landau-zener model, Journal of Physics A: Mathematical and General 19, 2061 (1986).